Myofascial Trigger points: an evidence-informed review

Myofascial Trigger Points: An Evidence-Informed Review
Jan Dommerholt, PT, MPS, FAAPM
Carel Bron, PT
Jo Franssen, PT
Received the 2006 OPTP Award for Excellence in a Published Review of the Literature
Abstract: This article provides a best evidence-informed review of the current scientific understanding of myofascial trigger points with regard to their etiology, pathophysiology, and
clinical implications. Evidence-informed manual therapy integrates the best available scientific evidence with individual clinicians’ judgments, expertise, and clinical decision-making.
After a brief historical review, the clinical aspects of myofascial trigger points, the interrater
reliability for identifying myofascial trigger points, and several characteristic features are
discussed, including the taut band, local twitch response, and referred pain patterns. The
etiology of myofascial trigger points is discussed with a detailed and comprehensive review of
the most common mechanisms, including low-level muscle contractions, uneven intramuscular pressure distribution, direct trauma, unaccustomed eccentric contractions, eccentric
contractions in unconditioned muscle, and maximal or sub-maximal concentric contractions.
Many current scientific studies are included and provide support for considering myofascial
trigger points in the clinical decision-making process. The article concludes with a summary
of frequently encountered precipitating and perpetuating mechanical, nutritional, metabolic,
and psychological factors relevant for physical therapy practice. Current scientific evidence
strongly supports that awareness and working knowledge of muscle dysfunction and in particular myofascial trigger points should be incorporated into manual physical therapy practice
consistent with the guidelines for clinical practice developed by the International Federation
of Orthopaedic Manipulative Therapists. While there are still many unanswered questions in
explaining the etiology of myofascial trigger points, this article provides manual therapists
with an up-to-date evidence-informed review of the current scientific knowledge.
K e y Wo rd s : M y o f a s c i a l P a i n S y n d r o m e , Tr i g g e r P o i n t s , M y o f a s c i a l , E t i o l o g y,
Pathophysiology
D
uring the past few decades, myofascial trigger points
(MTrPs) and myofascial pain syndrome (MPS) have
received much attention in the scientific and clinical
literature. Researchers worldwide are investigating
various aspects of MTrPs, including their specific etiology, pathophysiology, histology, referred pain patterns,
and clinical applications. Guidelines developed by the
International Federation of Orthopaedic Manipulative
Address all correspondence and request for reprints to:
Jan Dommerholt
Bethesda Physiocare, Inc.
7830 Old Georgetown Road, Suite C-15
Bethesda, MD 20814-2440
[email protected]
The Journal of Manual & Manipulative Therapy
Vol. 14 No. 4 (2006), 203 - 221
Therapists (IFOMT) confirm the importance of muscle
dysfunction for orthopedic manual therapy clinical practice. The IFOMT has defined orthopedic manual therapy
as “a specialized area of physiotherapy/physical therapy
for the management of neuromusculoskeletal conditions, based on clinical reasoning, using highly specific
treatment approaches including manual techniques and
therapeutic exercises.” The educational standards of
IFOMT require that skills will be demonstrated in—among
others—“analysis and specific tests for functional status
of the muscular system,” “a high level of skill in other
manual and physical therapy techniques required to
mobilize the articular, muscular or neural systems,” and
“knowledge of various manipulative therapy approaches
as practiced within physical therapy, medicine, osteopathy
and chiropractic”1.
Myofascial Trigger Points: An Evidence-Informed Review / 203
However, articles about muscle dysfunction in the
manual therapy literature are sparse and they generally
focus on muscle injury and muscle repair mechanisms2 or
on muscle recruitment3. Until very recently, the current
scientific knowledge and clinical implications of MTrPs
were rarely included4-7. It appears that orthopedic manual
therapists have not paid much attention to the pathophysiology and clinical manifestations of MTrPs. Manual
therapy educational programs in the US seem to reflect
this orientation and tend to place a strong emphasis on
joint dysfunction, mobilizations, and manipulations with
only about 10%-15% of classroom education devoted to
muscle pain and muscle dysfunction. This review of the MTrP literature is based on
current best scientific evidence. The field of manual
therapy has joined other medical disciplines by embracing evidence-based medicine, which proposes that the
results of scientific research need to be integrated into
clinical practice 8. Evidence-based medicine has been
defined as “the conscientious, explicit, and judicious
use of current best evidence in making decisions about
the care of individual patients”9,10. Within the evidencebased medicine paradigm, evidence is not restricted to
randomized controlled trials, systematic reviews, and
meta-analyses, although this restricted view seems to be
prevalent in the medical and physical therapy literature.
Sackett et al9,10 emphasized that external clinical evidence
can inform but not replace individual clinical expertise.
Clinical expertise determines whether external clinical
evidence applies to an individual patient, and if so, how
it should be integrated into clinical decision-making.
Pencheon11 shared this perspective and suggested that
high-quality healthcare is about combining “wisdom
produced by years of experience” with “evidence produced
by generalizable research” in “ways with which patients
are happy.” He suggested shifting from evidence-based
to evidence-informed medicine, where clinical decisionmaking is informed by research evidence but not driven
by it and always includes knowledge from experience.
Evidence-informed manual therapy involves integrating the best available external scientific evidence with
individual clinicians’ judgments, expertise, and clinical decision-making12. The purpose of this article is to
provide a best evidence-informed review of the current
scientific understanding of MTrPs, including the etiology,
pathophysiology, and clinical implications, against the
background of extensive clinical experience.
Brief Historical Review
While Dr. Janet Travell (1901-1997) is generally credited for bringing MTrPs to the attention of health care
providers, MTrPs have been described and rediscovered for
several centuries by various clinicians and researchers13,14.
As far back as the 16th century, de Baillou (1538-1616),
as cited by Ruhmann, described what is now known as
myofascial pain syndrome (MPS) 15. MPS is defined as
204 / The Journal of Manual & Manipulative Therapy, 2006
the “sensory, motor, and autonomic symptoms caused
by MTrPs” and has become a recognized medical diagnosis among pain specialists16,17. In 1816, British physician Balfour, as cited by Stockman, described “nodular
tumors and thickenings which were painful to the touch,
and from which pains shot to neighboring parts”18. In
1898, the German physician Strauss discussed “small,
tender and apple-sized nodules and painful, pencil-sized
to little-finger-sized palpable bands”19. The first trigger
point manual was published in 1931 in Germany nearly
a decade before Travell became interested in MTrPs20.
While these early descriptions may appear a bit archaic
and unusual—for example, in clinical practice one does
not encounter “apple-sized nodules” —these and other
historic papers did illustrate the basic features of MTrPs
quite accurately14.
In the late 1930s, Travell, who at that time was a
cardiologist and medical researcher, became particularly
interested in muscle pain following the publication of
several articles on referred pain21. Kellgren’s descriptions
of referred pain patterns of many muscles and spinal
ligaments after injecting these tissues with hypertonic
saline22-25 eventually moved Travell to shift her medical
career from cardiology to musculoskeletal pain. During
the 1940s, she published several articles on injection
techniques of MTrPs 26-28. In 1952, she described the
myofascial genesis of pain with detailed referred pain
patterns for 32 muscles29. Other clinicians also became
interested in MTrPs. European physicians Lief and Chaitow
developed a treatment method, which they referred to as
“neuromuscular technique”30. German physician Gutstein
described the characteristics of MTrPs and effective manual
therapy treatments in several papers under the names
of Gutstein, Gutstein-Good, and Good31-34. In Australia,
Kelly produced a series of articles about fibrositis, which
paralleled Travel’s writings35-38.
In the US, chiropractors Nimmo and Vannerson 39
described muscular “noxious generative points,” which
were thought to produce nerve impulses and eventually
result in “vasoconstriction, ischaemia, hypoxia, pain, and
cellular degeneration.” Later in his career, Nimmo adopted
the term “trigger point” after having been introduced
to Travell’s writings. Nimmo maintained that hypertonic
muscles are always painful to pressure, a statement that
later became known as “Nimmo’s law.” Like Travell,
Nimmo described distinctive referred pain patterns and
recommended releasing these dysfunctional points by
applying the proper degree of manual pressure. Nimmo’s
“receptor-tonus control method” continues to be popular
among chiropractic physicians39,40. According to a 1993
report by the National Board of Chiropractic Economics,
over 40% of chiropractors in the US frequently apply
Nimmo’s techniques41. Two spin-offs of Nimmo’s work
are St. John Neuromuscular Therapy (NMT) method and
NMT American version, which have become particularly
popular among massage therapists30.
In 1966, Travell founded the North American Academy
of Manipulative Medicine, together with Dr. John Mennell,
who also published several articles about MTrPs 42,43.
Throughout her career Travell promoted integrating
myofascial treatments with articular treatments16. One
of her earlier papers described a technique for reducing sacroiliac displacement44. However, Travell, as cited
by Paris45, maintained the opinion that manipulations
were the exclusive domain of physicians and she rejected membership in the North American Academy of
Manipulative Medicine by physical therapists.
In the early 1960s, Dr. David Simons was introduced
to Travell and her work, which became the start of a
fruitful collaboration eventually resulting in several publications, including the Trigger Point Manuals, consisting of a 1983 first volume (upper half of the body) and
a 1992 second volume (lower half of the body)46,47. The
first volume has since been revised and updated and a
second edition was released in 199916. The Trigger Point
Manuals are the most comprehensive review of nearly
150 muscle referred-pain patterns based on Travell’s
clinical observations, and they include an extensive
review of the scientific basis of MTrPs. Both volumes
have been translated into several foreign languages,
including Russian, German, French, Italian, Japanese,
and Spanish. Several other clinicians worldwide have
also published their own trigger point manuals48-54.
observed significant lowering of the pain threshold over
active MTrPs when measured by electrical stimulation,
not only in the muscular tissue but also in the overlying
cutaneous and subcutaneous tissues. In contrast, with
latent MTrPs, the sensory changes did not involve the
cutaneous and subcutaneous tissues57-59. Autonomic aspects
of MTrPs may include, among others, vasoconstriction,
vasodilatation, lacrimation, and piloerection16,60-63.
A detailed clinical history, examination of movement
patterns, and consideration of muscle referred-pain patterns assist clinicians in determining which muscles
may harbor clinically relevant MTrPs64. Muscle pain is
perceived as aching and poorly localized. There are no
laboratory or imaging tests available that can confirm the
presence of MTrPs. Myofascial trigger points are identified through either a flat palpation technique (Figure 1)
in which a clinician applies finger or thumb pressure
to muscle against underlying bone tissue, or a pincer
palpation technique (Figure 2) in which a particular
muscle is palpated between the clinician’s fingers. Clinical Aspects of
Myofascial Trigger Points
An MTrP is described as “a hyperirritable spot in
skeletal muscle that is associated with a hypersensitive
palpable nodule in a taut band”16. Myofascial trigger points
are classified into active and latent trigger points16. An
active MTrP is a symptom-producing MTrP and can trigger
local or referred pain or other paraesthesiae. A latent
MTrP does not trigger pain without being stimulated.
Myofascial trigger points are the hallmark characteristics of MPS and feature motor, sensory, and autonomic
components. Motor aspects of active and latent MTrPs
may include disturbed motor function, muscle weakness as a result of motor inhibition, muscle stiffness,
and restricted range of motion55,56. Sensory aspects may
include local tenderness, referral of pain to a distant
site, and peripheral and central sensitization. Peripheral
sensitization can be described as a reduction in threshold
and an increase in responsiveness of the peripheral ends
of nociceptors, while central sensitization is an increase
in the excitability of neurons within the central nervous
system. Signs of peripheral and central sensitization are
allodynia (pain due to a stimulus that does not normally
provoke pain) and hyperalgesia (an increased response
to a stimulus that is normally painful). Both active and
latent MTrPs are painful on compression. Vecchiet et al5759
described specific sensory changes over MTrPs. They
Fig. 1: Flat palpation
Fig. 2: Pincer palpation
Myofascial Trigger Points: An Evidence-Informed Review / 205
By definition, MTrPs are located within a taut band of
contractured muscle fibers (Figure 3), and palpating for
MTrPs starts with identifying this taut band by palpating
perpendicular to the fiber direction. Once the taut band
is located, the clinician moves along the taut band to
Fig. 3: Palpation of a trigger point within a taut band
(reproduced with permission from Weisskircher H-W. Head
Pains Due to Myofascial Trigger Points. CD-ROM, www.
trigger-point.com, 1997)
find a discrete area of intense pain and hardness.
Two studies have reported good overall interrater
reliability for identifying taut bands, MTrPs, referred pain,
and local twitch responses 65,66. The minimum criteria
that must be satisfied in order to distinguish an MTrP
from any other tender area in muscle are a taut band
and a tender point in that taut band65. Although Janda
maintained that systematic palpation can differentiate
between myofascial taut bands and general muscle spasms,
electromyography is the gold standard to differentiate taut
bands from contracted muscle fibers 67,68. Spasms can be
defined as electromyographic (EMG) activity as the result
of increased neuromuscular tone of the entire muscle,
and they are the result of nerve-initiated contractions. A
taut band is an endogenous localized contracture within
the muscle without activation of the motor endplate69.
From a physiological perspective, the term “contracture”
is more appropriate then “contraction” when describing
chronic involuntary shortening of a muscle without
EMG activity. In clinical practice, surface EMG is used
in the diagnosis and management of MTrPs in addition
to manual examinations 67,70,71. Diagnostically, surface
EMG can assist in assessing muscle behavior during rest
and during functional tasks. Clinicians use the MTrP
referred pain patterns in determining which muscles
to examine with surface EMG. Muscles that harbor
MTrPs responsible for the patient’s pain complaint are
examined first. EMG assessments guide the clinician
with postural training, ergonomic interventions, and
muscle awareness training67.
The patient’s recognition of the elicited pain further
206 / The Journal of Manual & Manipulative Therapy, 2006
guides the clinician. The presence of a so-called local
twitch response (LTR), referred pain, or reproduction of
the person’s symptomatic pain increases the certainty
and specificity of the diagnosis of MPS. Local twitch
responses are spinal reflexes that appear to be unique
to MTrPs. They are characterized by a sudden contraction of muscle fibers within a taut band, when the taut
band is strummed manually or needled. The sudden
contractions can be observed visually, can be recorded
electromyographically, or can be visualized with diagnostic ultrasound 72. When an MTrP is needled with a
monopolar teflon-coated EMG needle, LTRs appear as
high-amplitude poly-phasic EMG discharges73-78.
In clinical practice, there is no benefit in using
needle EMG or sonography, and its utility is limited to
research studies. For example, Audette et al79 established
that in 61.5% of active MTrPs in the trapezius and levator
scapulae muscles, dry needling an active MTrP elicited
an LTR in the same muscle on the opposite side of the
body. Needling of latent MTrPs resulted in unilateral
LTRs only. In this study, LTRs were used to research
the nature of active versus latent MTrPs. Studies have
shown that clinical outcomes are significantly improved
when LTRs are elicited in the treatment of patients with
dry needling or injection therapy 74,80,81. The taut band,
MTrP, and LTR (Figure 4) are objective criteria, identified
solely by palpation, that do not require a verbal response
from the patient82. Active MTrPs refer pain usually to a distant site.
The referred pain patterns (Figure 5) are not necessarily
Fig. 4: Local twitch response in a rabbit trigger spot.
Local twitch responses are elicited only when the needle
is placed accurately within the trigger spot. Moving as
little as 0.5 cm away from the trigger spot virtually
eliminates the local twitch response (reproduced with
permission from Hong C-Z, Torigoe Y. Electrophysiological
characteristics of localized twitch responses in responsive
taut bands of rabbit skeletal muscle. J Musculoskeletal
Pain 1994;2:17-43)
Fig. 5: MTrP referred pain patterns (reproduced with permission from MEDICLIP, Manual Medicine 1 & 2, Version
1.0a, 1997, Williams & Wilkins)
restricted to single segmental pathways or to peripheral nerve distributions. Although typical referred pain
patterns have been established, there is considerable
variation between patients16,48.
Usually, the pain in reference zones is described as
“deep tissue pain” of a dull and aching nature. Occasionally, patients may report burning or tingling sensations,
especially in superficial muscles such as the platysma
muscle 83,84. By mechanically stimulating active MTrPs,
patients may report the reproduction of their pain, either
immediately or after a 10-15 second delay. Normally,
skeletal muscle nociceptors require high intensities of
stimulation and they do not respond to moderate local
pressure, contractions, or muscle stretches85. However,
MTrPs cause persistent noxious stimulation, which results
in increasing the number and size of the receptive fields to
which a single dorsal horn nociceptive neuron responds,
and the experience of spontaneous and referred pain 86.
Several recent studies have determined previously unrecorded referred pain patterns of different muscles and
MTrPs87-90. Referred pain is not specific to MPS but it is
relatively easy to elicit over MTrPs91. Normal muscle tissue
and other body tissues, including the skin, zygapophyseal
joints, or internal organs, may also refer pain to distant
regions with mechanical pressure, making referred pain
elicited by stimulation of a tender location a nonspecific
finding84,92-95. Gibson et al96 found that referred pain is actually easier to elicit in tendon-bone junctions and tendon
than in the muscle belly. However, after exposing the
muscle to eccentric exercise, significantly higher referred
pain frequency and enlarged pain areas were found at the
muscle belly and the tendon-bone junction sites following
injection with hypotonic saline. The authors suggested
that central sensitization may explain the referred pain
frequency and enlarged pain areas97.
While a survey of members of the American Pain
Society showed general agreement that MTrPs and MPS
exist as distinct clinical entities, MPS continues to be
one of the most commonly missed diagnoses 17,98. In a
recent study of 110 adults with low back pain, myofascial
pain was the most common finding affecting 95.5% of
patients, even though myofascial pain was poorly defined
as muscle pain in the paraspinal muscles, piriformis, or
tensor fasciae latae99. A study of adults with frequent migraine headaches diagnosed according to the International
Headache Society criteria showed that 94% of the patients
reported migrainous pain with manual stimulation of
cervical and temporal MTrPs, compared with only 29% of
controls100,101. In 30% of the migraine group, palpation of
MTrPs elicited a “full-blown migraine attack that required
abortive treatment.” The researchers found a positive
relationship between the number of MTrPs and the frequency of migraine attacks and duration of the illness100.
Several studies have confirmed that MTrPs are common
not only in persons attending pain management clinics
but also in those seeking help through internal medicine
and dentistry102-107. In fact, MTrPs have been identified
with nearly every musculoskeletal pain problem, including radiculopathies104, joint dysfunction108, disk pathology109, tendonitis110, craniomandibular dysfunction111-113,
migraines100,114, tension-type headaches7,87, carpal tunnel
syndrome115, computer-related disorders116, whiplash-associated disorders60,117, spinal dysfunction118, and pelvic
pain and other urologic syndromes119-122. Myofascial trigger
points are associated with many other pain syndromes123,
including, for example, post-herpetic neuralgia124,125, complex
regional pain syndrome126,127, nocturnal cramps128, phantom
pain129,130, and other relatively uncommon diagnoses such
as Barré-Liéou syndrome131 and neurogenic pruritus132. A
recent study suggested that there might be a relationship
between MTrPs in the upper trapezius muscle and cervical
spine dysfunction at the C3 and C4 vertebrae, although
a cause-and-effect relationship was not established in
this correlational study133. Another study described that
persons with mechanical neck pain had significantly
more clinically relevant MTrPs in the upper trapezius,
sternocleidomastoid, levator scapulae, and suboccipital
muscles as compared to healthy controls5.
Etiology of MTrPs
Several possible mechanisms can lead to the development of MTrPs, including low-level muscle contractions, uneven intramuscular pressure distribution, direct
trauma, unaccustomed eccentric contractions, eccentric
contractions in unconditioned muscle, and maximal or
submaximal concentric contractions. Low-level muscle contractions
Of particular interest in the etiology of MTrPs are
low-level muscle exertions and the so-called Cinderella
Myofascial Trigger Points: An Evidence-Informed Review / 207
Hypothesis developed by Hägg in 1988134. The Cinderella
Hypothesis postulates that occupational myalgia is caused
by selective overloading of the earliest recruited and
last de-recruited motor units according to the ordered
recruitment principle or Henneman’s “size principle”134,135.
Smaller motor units are recruited before and de-recruited
after larger ones; as a result, the smaller type 1 fibers are
continuously activated during prolonged motor tasks135.
According to the Cinderella Hypothesis, muscular force
generated at sub-maximal levels during sustained muscle
contractions engages only a fraction of the motor units
available without the normally occurring substitution of
motor units during higher force contractions, which in
turn can result in metabolically overloaded motor units,
prone to loss of cellular Ca2+-homeostasis, subsequent
activation of autogenic destructive processes, and muscle
pain136,137. The other pillar of the Cinderella Hypothesis is
the finding of an excess of ragged red fibers in myalgic
patients136. Indeed, several researchers have demonstrated
the presence of ragged red fibers and moth-eaten fibers
in subjects with myalgia, which are indications of structural damage to the cell membrane and mitochondria
and a change in the distribution of mitochondria or the
sarcotubular system respectively138-142.
There is growing evidence that low-level static muscle
contractions or exertions can result in degeneration of
muscle fibers143. Gissell144,145 has shown that low-level
exertions can result in an increase of Ca 2+-release in
skeletal muscle cells, muscle membrane damage due to
leakage of the intracellular enzyme lactate dehydrogenase,
structural damage, energy depletion, and myalgia. Lowlevel muscle stimulation can also lead to the release of
interleukin 6 (IL-6) and other cytokines146,147. Several studies have confirmed the Cinderella Hypothesis and support the idea that in low-level static
exertions, muscle fiber recruitment patterns tend to
be stereotypical with continuous activation of smaller
type 1 fibers during prolonged motor tasks 148-152. As
Hägg indicated, the continuous activity and metabolic
overload of certain motor units does not occur in all
subjects 136. The Cinderella Hypothesis was recently
applied to the development of MTrPs116. In a well-designed study, Treasters et al116 established that sustained
low-level muscle contractions during continuous typing
for as little as 30 minutes commonly resulted in the
formation of MTrPs. They suggested that MTrPs might
provide a useful explanation for muscle pain and injury
that can occur from low-level static exertions116. Myofascial trigger points are common in office workers,
musicians, dentists, and other occupational groups
exposed to low-level muscle exertions153. Chen et al154
also suggested that low-level muscle exertions can
lead to sensitization and development of MTrPs. Forty
piano students showed significantly reduced pressure
thresholds over latent MTrPs after only 20 minutes of
continuous piano playing154.
208 / The Journal of Manual & Manipulative Therapy, 2006
Intramuscular pressure distribution
Otten155 has suggested that circulatory disturbances
secondary to increased intramuscular pressure may also
lead to the development of myalgia. Based on mathematical modeling applied to a frog gastrocnemius muscle,
Otten confirmed that during static low-level muscle
contractions, capillary pressures increase dramatically
especially near the muscle insertions (Figure 6). In other
words, during low-level exertions, the intramuscular
Fig. 6: Intramuscular pressure distribution in the gastrocnemius muscle of the toad (reproduced with permission
from E. Otten, 2006)
pressure near the muscle insertions might increase
rapidly, leading to excessive capillary pressure, decreased
circulation, and localized hypoxia and ischaemia155. With
higher level contractions in between 10% and 20% of
maximum voluntary effort, the intramuscular pressure
increases also in the muscle belly 156,157. According to
Otten, the increased pressure gradients during low-level
exertions may contribute to the development of pain at
the musculotendinous junctions and eventually to the
formation of MTrPs (personal communication, 2005).
In 1999, Simons introduced the concept of “attachment trigger points” to explain pain at the musculotendinous junctions in persons with MTrPs, based on the
assumption that taut bands would generate sufficient
sustained force to induce localized enthesopathies16,158.
More recently, Simons concluded that there is no convincing evidence that the tension generated in shortened
sarcomeres in a muscle belly would indeed be able to
generate passive or resting force throughout an entire
taut band resulting in enthesopathies, even though
there may be certain muscles or conditions where this
could occur (personal communication, 2005). To the
contrary, force generated by individual motor units is
always transmitted laterally to the muscle’s connective
tissue matrix, involving at least two protein complexes
containing vinculin and dystrophin, respectively159. There
is also considerable evidence that the assumption that
muscle fibers pass from tendon to tendon is without
basis160. Trotter160 has demonstrated that skeletal muscle
is comprised of in-series fibers. In other words, there is
evidence that a single muscle fiber does not run from
tendon to tendon. The majority of fibers are in series
with inactive fibers, which makes it even more unlikely
that the whole muscle length-tension properties would
be dictated by the shortest contractured fibers in the
muscle161.
In addition, it is important to consider the mechanical
and functional differences between fast and slow motor
units162,163. Slow motor units are always stiffer than fast
units, although fast units can produce more force. If
there were any transmission of force along the muscle
fiber, as Simons initially suggested, fast fibers would be
better suited to accomplish this. Yet, fast motor units
have larger series of elastic elements, which would
absorb most of the force displacement164,165. Fast fibers
show a progressive decrease in cross-sectional area and
end in a point within the muscle fascicle, making force
transmission even more unlikely163. Fast fibers rely on
transmitting a substantial proportion of their force to
the endomysium, transverse cytoskeleton, and adjacent muscle fibers162,163. In summary, the development
of so-called “attachment trigger points” as a result of
increased tension by contractured sarcomeres in MTrPs
is not clear and more research is needed to explain the
clinical observation that MTrPs appear to be linked to
pain at the musculotendinous junction. The increased
tension in the muscle belly is likely to dissipate across
brief sections of the taut band on both sides of the MTrP
and laterally through the transverse cytoskeleton166-168.
Instead, Otten’s model of increased intramuscular pressure,
decreased circulation, localized hypoxia, and ischaemia at
the muscle insertions provides an alternative model for
the clinically observed pain near the musculotendinous
junction and osseous insertions in persons with MTrPs,
even though the model does not explain why taut bands
are commonly present155.
Direct trauma
There is general agreement that acute muscle overload can activate MTrPs, although systematic studies
are lacking169. For example, people involved in whiplash
injuries commonly experience prolonged muscle pain
and dysfunction170-173. In a retrospective review, Schuller et al174 found that 80% of 1096 subjects involved in
low-velocity collisions demonstrated evidence of muscle
pain with myogeloses among the most common find-
ings. Although Schuller et al 174 did not define these
myogeloses, Simons has suggested that a myogelosis
describes the same clinical entity as an MTrP175. Baker117
reported that the splenius capitis, semispinalis capitis,
and sternocleidomastoid muscles developed symptomatic
MTrPs in 77%, 62%, and 52% of 52 whiplash patients,
respectively. In a retrospective review of 54 consecutive
chronic whiplash patients, Gerwin and Dommerholt 176
reported that clinically relevant MTrPs were found in
every patient, with the trapezius muscle involved most
often. Following treatment emphasizing the inactivation
of MTrPs and restoration of normal muscle length, approximately 80% of patients experienced little or no pain,
even though the average time following the initiating
injury was 2.5 years at the beginning of the treatment
regimen. All patients had been seen previously by other
physicians and physical therapists who apparently had
not considered MTrPs in their thought process and
clinical management176. Fernández-de-las-Peñas et al177,178
confirmed that inactivation of MTrPs should be included
in the management of persons suffering from whiplashassociated disorders. In their research-based treatment
protocol, the combination of cervical and thoracic spine
manipulations with MTrP treatments proved superior
to more conventional physical therapy consisting of
massage, ultrasound, home exercises, and low-energy
high-frequency pulsed electromagnetic therapy177.
Direct trauma may create a vicious cycle of events
wherein damage to the sarcoplasmic reticulum or the
muscle cell membrane may lead to an increase of the
calcium concentration, a subsequent activation of actin
and myosin, a relative shortage of adenosine triphosphate
(ATP), and an impaired calcium pump, which in turn
will increase the intracellular calcium concentration
even more, completing the cycle. The calcium pump is
responsible for returning intracellular Ca2+ to the sarcoplasmic reticulum against a concentration gradient,
which requires a functional energy supply. Simons and
Travell179 considered this sequence in the development
of the so-called “energy crisis hypothesis” introduced
in 1981. Sensory and motor system dysfunction have
been shown to develop rapidly after injury and actually
may persist in those who develop chronic muscle pain
and in individuals who have recovered or continue to
have persistent mild symptoms172,180. Scott et al181 determined that individuals with chronic whiplash pain
develop more widespread hypersensitivity to mechanical
pressure and thermal stimuli than those with chronic
idiopathic neck pain. Myofascial trigger points are a
likely source of ongoing peripheral nociceptive input,
and they contribute to both peripheral and central
sensitization, which may explain the observation of
widespread allodynia and hypersensitivity60,62,63. In addition to being caused by whiplash injury, acute muscle
overload can occur with direct impact, lifting injuries,
sports performance, etc.182.
Myofascial Trigger Points: An Evidence-Informed Review / 209
Eccentric and (sub)maximal concentric contractions
Many patients report the onset of pain and activation
of MTrPs following either acute, repetitive, or chronic
muscle overload183. Gerwin et al184 suggested that likely
mechanisms relevant for the development of MTrPs
included either unaccustomed eccentric exercise, eccentric exercise in unconditioned muscle, or maximal
or sub-maximal concentric exercise. A brief review of
pertinent aspects of exercise follows, preceding linking
this body of research to current MTrP research.
Eccentric exercise is associated with myalgia, muscle
weakness, and destruction of muscle fibers, partially
because eccentric contractions cause an irregular and
uneven lengthening of muscle fibers185-187. Muscle soreness and pain occur because of local ultra-structural
damage, the release of sensitizing algogenic substances,
and the subsequent onset of peripheral and central
sensitization85,188-190. Muscle damage occurs at the cytoskeletal level and frequently involves disorganization of
the A-band, streaming of the Z-band, and disruption of
cytoskeletal proteins, such as titin, nebulin, and desmin,
even after very short bouts of eccentric exercise186,189-194.
Loss of desmin can occur within 5 minutes of eccentric
loading, even in muscles that routinely contract eccentrically during functional activities, but does not occur
after isometric or concentric contractions193,195. Lieber
and Fridén193 suggested that the rapid loss of desmin
might indicate a type of enzymatic hydrolysis or protein
phosphorylation as a likely mechanism.
One of the consequences of muscle damage is muscle
weakness196-198. Furthermore, concentric and eccentric
contractions are linked to contraction-induced capillary constrictions, impaired blood flow, hypoperfusion,
ischaemia, and hypoxia, which in turn contribute to
the development of more muscle damage, a local acidic
milieu, and an excessive release of protons (H+), potassium
(K+), calcitonin-gene-related-peptide (CGRP), bradykinin
(BK), and substance P (SP), and sensitization of muscle
nociceptors 184,188. There are striking similarities with
the chemical environment of active MTrPs established
with microdialysis, suggesting an overlap between the
research on eccentric exercise and MTrP research184,199.
However, at this time, it is premature to conclude that
there is solid evidence that eccentric and sub-maximal
concentric exercise are absolute precursors to the development of MTrPs. In support of this hypothesized
causal relation, Itoh et al200 demonstrated in a recent
study that eccentric exercise can lead to the formation
of taut and tender ropy bands in exercised muscle, and
they hypothesized that eccentric exercise may indeed be
a useful model for the development of MTrPs.
Eccentric and concentric exercise and MTrPs have
been associated with localized hypoxia, which appears
to be one of the most important precursors for the
development of MTrPs201. As mentioned, hypoxia leads
to the release of multiple algogenic substances. In this
210 / The Journal of Manual & Manipulative Therapy, 2006
context, recent research by Shah et al199 at the US National Institutes of Health is particularly relevant. Shah
et al analyzed the chemical milieu of latent and active
MTrPs and normal muscles. They found significantly increased concentrations of BK, CGRP, SP, tumor necrosis
factor-α (TNF-α), interleukin-1β (IL-1β), serotonin, and
norepinephrine in the immediate milieu of active MTrPs
only199. These substances are well-known stimulants for
various muscle nociceptors and bind to specific receptor
molecules of the nerve endings, including the so-called
purinergic and vanilloid receptors85,202.
Muscle nociceptors are dynamic structures whose
receptors can change depending on the local tissue
environment. When a muscle is damaged, it releases
ATP, which stimulates purinergic receptors, which are
sensitive to ATP, adenosine diphosphate, and adenosine.
They bind ATP, stimulate muscle nociceptors, and cause
pain. Vanilloid receptors are sensitive to heat and respond
to an increase in H+-concentration, which is especially
relevant under conditions with a lowered pH, such as
ischaemia, inflammation, or prolonged and exhaustive
muscle contractions85. Shah et al199 determined that the
pH at active MTrP sites is significantly lower than at
latent MTrP sites. A lowered pH can initiate and maintain muscle pain and mechanical hyperalgesia through
activation of acid-sensing ion channels203,204. Neuroplastic
changes in the central nervous system facilitate mechanical hyperalgesia even after the nociceptive input
has been terminated (central sensitization) 203,204. Any
noxious stimulus sufficient to cause nociceptor activation causes bursts of SP and CGRP to be released into
the muscle, which have a significant effect on the local
biochemical milieu and microcirculation by stimulating
“feed-forward” neurogenic inflammation. Neurogenic
inflammation can be described as a continuous cycle of
increasing production of inflammatory mediators and
neuropeptides and an increasing barrage of nociceptive
input into wide dynamic-range neurons in the spinal
cord dorsal horn184.
The Integrated Trigger Point Hypothesis
The integrated trigger point hypothesis (Figure 7)
has evolved since its first introduction as the “energy
crisis hypothesis” in 1981. It is based on a combination of
electrodiagnostic and histopathological evidence179,183.
Already in 1957, Weeks and Travell205 had published
a report that outlined a characteristic electrical activity of an MTrP. It was not until 1993 that Hubbard et
al 206 confirmed that this EMG discharge consists of
low-amplitude discharges in the order of 10-50 µV and
intermittent high-amplitude discharges (up to 500 µV)
in painful MTrPs. Initially, the electrical activity was
termed “spontaneous electrical activity” (SEA) and
thought to be related to dysfunctional muscle spindles206.
Best available evidence now suggests that the SEA is in
fact endplate noise (EPN), which is found much more
Fig. 7: The integrated trigger point hypothesis.
Ach- acetylcholine; AchE- acetylcholinesterase; AchRacetylcholine receptor
commonly in the endplate zone near MTrPs than in an
endplate zone outside MTrPs207-209. The electrical discharges
occur with frequencies that are 10-1,000 times that of
normal endplate potentials, and they have been found in
humans, rabbits, and recently even in horses209,210. The
discharges are most likely the result of an abnormally
excessive release of acetylcholine (ACh) and indicative
of dysfunctional motor endplates, contrary to the commonly accepted notion among electromyographers that
endplate noise arises from normal motor endplates183.
The effectiveness of botulinum toxin in the treatment
of MTrPs provides indirect evidence of the presence of
excessive ACh211. Botulinum toxin (BoTox) is a neurotoxin
that blocks the release of ACh from presynaptic cholinergic nerve endings. A recent study in mice demonstrated
that the administration of botulinum toxin resulted in a
complete functional repair of dysfunctional endplates212.
There is some early evidence that muscle stretching and
hypertonicity may also enhance the excessive release of
ACh213,214. Tension on the integrins in the presynaptic
membrane at the motor nerve terminal is hypothesized
to mechanically trigger an ACh release that does not
require Ca 2+ 213-215. Integrins are receptor proteins in the
cell membrane involved in attaching individual cells to
the extracellular matrix.
Excessive ACh affects voltage-gated sodium channels of the sarcoplasmic reticulum and increases the
intracellular calcium levels, which triggers sustained
muscle contractures. It is conceivable that in MTrPs,
myosin filaments literally get stuck in the Z-band of
the sarcomere. During sarcomere contractions, titin
filaments are folded into a gel-like structure at the Zband. In MTrPs, the gel-like titin may prevent the myosin
filaments from detaching. The myosin filaments may
actually damage the regular motor assembly and prevent
the sarcomere from restoring its resting length216. Muscle
contractures are also maintained because of the relative
shortage of ATP in an MTrP, as ATP is required to break
the cross-bridges between actin and myosin filaments.
The question remains whether sustained contractures
require an increase of oxygen availability. At the same time, the shortened sarcomeres compromise the local circulation causing ischaemia. Studies of
oxygen saturation levels have demonstrated severe hypoxia
in MTrPs201. Hypoxia leads to the release of sensitizing
substances and activates muscle nociceptors as reviewed
above. The combined decreased energy supply and possible increased metabolic demand would also explain
the common finding of abnormal mitochondria in the
nerve terminal and the previously mentioned ragged red
fibers. In mice, the onset of hypoxia led to an immediate
increased ACh release at the motor endplate217.
The combined high-intensity mechanical and chemical stimuli may cause activation and sensitization of
the peripheral nerve endings and autonomic nerves,
activate second order neurons including so-called “sleeping” receptors, cause central sensitization, and lead to
the formation of new receptive fields, referred pain, a
long-lasting increase in the excitability of nociceptors,
and a more generalized hyperalgesia beyond the initial
nociceptive area. An expansion of a receptive field means
that a dorsal horn neuron receives information from
areas it has not received information from previously218.
Sensitization of peripheral nerve endings can also cause
pain through SP activating the neurokin-1 receptors
and glutamate activating N-methyl-D-aspartate receptors, which opens post-synaptic channels through which
Ca2+ ions can enter the dorsal horn and activate many
enzymes involved in the sensitization85.
Several histological studies offer further support for
the integrated trigger point hypothesis. In 1976, Simons
and Stolov published the first biopsy study of MTrPs
in a canine muscle and reported multiple contraction
knots in various individual muscle fibers (Figure 8)219.
The knots featured a combination of severely shortened
sarcomeres in the center and lengthened sarcomeres
outside the immediate MTrP region219.
Reitinger et al220 reported pathologic alterations of
the mitochondria as well as increased width of A-bands
and decreased width of I-bands in muscle sarcomeres of
MTrPs in the gluteus medius muscle. Windisch et al221
determined similar alterations in a post-mortem histological study of MTrPs completed within 24 hours of time
of death. Mense et al 222 studied the effects of electrically
induced muscle contractions and a cholinesterase blocker
on muscles with experimentally induced contraction knots
and found evidence of localized contractions, torn fibers,
and longitudinal stripes. Pongratz and Spath223, 224 demonstrated evidence of a contraction disk in a region of an
MTrP using light microscopy. New MTrP histopathological
studies are currently being conducted at the Friedrich
Myofascial Trigger Points: An Evidence-Informed Review / 211
Fig. 8: Longitudinal section of a contraction knot in a canine
gracilis muscle (reproduced with permission from: Simons
DG, Travell JG, Simons LS. Travell and Simons’ Myofascial
Pain and Dysfunction: The Trigger Point Manual. Vol. 1.
2nd ed. Baltimore, MD: Williams & Wilkins, 1999)
Baur Institute in Munich, Germany. Gariphianova 225
described pathological changes with biopsy studies of
MTrPs, including a decrease in quantity of mitochondria,
possibly indicating metabolic distress. Several older
histological studies are often quoted, but it is not clear
to what extent those findings are specific for MTrPs. In
1951, Glogowsky and Wallraff226 reported damaged fibril
structures. Fassbender227 observed degenerative changes
of the I-bands, in addition to capillary damage, a focal
accumulation of glycogen, and a disintegration of the
myofibrillar network.
There is growing evidence for the integrated trigger
point hypothesis with regard to the motor and sensory
aspects of MTrPs, but many questions remain about
the autonomic aspects. Several studies have shown
that MTrPs are influenced by the autonomic nervous
system. Exposing subjects with active MTrPs in the
upper trapezius muscles to stressful tasks consistently
increased the electrical activity in MTrPs in the upper
trapezius muscle but not in control points in the same
muscle, while autogenic relaxation was able to reverse
the effects228-231. The administration of the sympathetic
blocking agent phentolamine significantly reduced the
electrical activity of an MTrP228,232,233. The interactions
between the autonomic nervous system and MTrPs need
further investigation. Hubbard 228 maintained that the
autonomic features of MTrPs are evidence that MTrPs
may be dysfunctional muscle spindles. Gerwin et al 184
have suggested that the presence of alpha and beta
adrenergic receptors at the endplate provide a possible
mechanism for autonomic interaction. In a rodent,
stimulation of the alpha and beta adrenergic receptors
stimulated the release of ACh in the phrenic nerve234.
In a recent study, Ge et al61 provided for the first time
experimental evidence of sympathetic facilitation of me-
212 / The Journal of Manual & Manipulative Therapy, 2006
chanical sensitization of MTrPs, which they attributed to
a change in the local chemical milieu at the MTrPs due
to increased vasoconstriction, an increased sympathetic
release of noradrenaline, or an increased sensitivity to
noradrenaline. Another intriguing possibility is that the
cytokine interleukin-8 (IL-8) found in the immediate
milieu of active MTrPs may contribute to the autonomic
features of MTrP. IL-8 can induce mechanical hyper-nociception, which is inhibited by beta adrenergic receptor
antagonists235. Shah et al found significantly increased
levels of IL-8 in the immediate milieu of active MTrPs
(Shah, 2006, personal communication).
The findings of Shah et al199 mark a major milestone
in the understanding and acceptance of MTrPs and support
parts of the integrated trigger point hypothesis183. The
possible consequences of several of the chemicals present
in the immediate milieu of active MTrPs have been
explored by Gerwin et al184. As stated, Shah et al found
significantly increased concentrations of H+, BK, CGRP,
SP, TNF-α, IL-1β, serotonin, and norepinephrine in active
MTrPs only. There are many interactions between these
chemicals that all can contribute to the persistent nature
of MTrPs through various vicious feedback cycles236. For
example, BK is known to activate and sensitize muscle
nociceptors, which leads to inflammatory hyperalgesia,
an activation of high-threshold nociceptors associated
with C-fibers, and even an increased production of BK
itself. Furthermore, BK stimulates the release of TNF-α,
which activates the production of the interleukins IL-1β,
IL-6, and IL-8. Especially IL-8 can cause hyperalgesia
that is independent from prostaglandin mechanisms.
Via a positive feedback loop, IL-1β can also induce the
release of BK237. Release of BK, K+, H+, and cytokines
from injured muscle activates the muscle nociceptors,
thereby causing tenderness and pain184.
Calcitonin gene-related peptide can enhance the
release of ACh from the motor endplate and simultaneously decrease the effectiveness of acetylcholinesterase
(AChE) in the synaptic cleft, which decreases the removal
of ACh 238,239. Calcitonin gene-related peptide also upregulates the ACh-receptors (AChR) at the muscle and
thereby creates more docking stations for ACh. Miniature
endplate activity depends on the state of the AChR and
on the local concentration of ACh, which is the result
of ACh-release, reuptake, and breakdown by AChE. In
summary, increased concentrations of CGRP lead to a
release of more ACh, and increase the impact of ACh by
reducing AChE effectiveness and increasing AChR efficiency.
Miniature endplate potential frequency is increased as
a result of greater ACh effect. The observed lowered pH
has several implications as well. Not only does a lower
pH enhance the release of CGRP, it also contributes to a
further down-regulation of AChE. The multiple chemicals
and lowered pH found in active MTrPs can contribute
to the chronic nature of MTrPs, enhance the segmental
spread of nociceptive input into the dorsal horn of the
spinal cord, activate multiple receptive fields, and trigger
referred pain, allodynia, hypersensitivity, and peripheral
and central sensitization that are characteristic of active
myofascial MTrPs184. There is no other evidence-based
hypothesis that explains the phenomena of MTrPs in
as much detail and clarity as the expanded integrated
trigger point hypothesis (Figure 9).
Fig. 9: The expanded MTrP hypothesis (reproduced with
permission from: Gerwin RD, Dommerholt J, Shah J. An
expansion of Simons’ integrated hypothesis of trigger point
formation. Curr Pain Headache Rep 2004;8:468-475).
Ach- acetylcholine; AchE- acetylcholinesterase; AchRacetylcholine receptor; ATP- adenosine triphosphate;
SP- substance P; CGRP- calcitonin gene-related peptide;
MEPP- miniature endplate potential
Perpetuating Factors
There are several precipitating or perpetuating factors
that need to be identified and, if present, adequately
managed to successfully treat persons with chronic
myalgia. Even though several common perpetuating
factors are more or less outside the direct scope of
manual physical therapy, familiarity with these factors
is critical especially considering the development of
increasingly autonomous physical therapy practice.
Simons, Travell, and Simons 16 identified mechanical,
nutritional, metabolic, and psychological categories of
perpetuating factors. Mechanical factors are familiar to
manual therapists and include the commonly observed
forward head posture, structural leg length inequalities,
scoliosis, pelvic torsion, joint hypermobility, ergonomic
stressors, poor body mechanics, etc.16,102,116,240.
In recent review articles, Gerwin 241,242 provided a
comprehensive update with an emphasis on non-structural perpetuating factors. Management of these factors
usually requires an interdisciplinary approach, including
medical and psychological intervention 64,82. Common
nutritional deficiencies or insufficiencies involve vitamin
B1, B6, B12, folic acid, vitamin C, vitamin D, iron,
magnesium, and zinc, among others. The term “insufficiency” is used to indicate levels in the lower range of
normal, such as those associated with biochemical or
metabolic abnormalities or with subtle clinical signs and
symptoms. Nutritional or metabolic insufficiencies are
frequently overlooked and not necessarily considered
clinically relevant by physicians unfamiliar with MTrPs
and chronic pain conditions. Yet any inadequacy that
interferes with the energy supply of muscle is likely to
aggravate MTrPs242. The most common deficiencies and
insufficiencies will be reviewed briefly. Vitamin B12 deficiencies are rather common and
may affect as many as 15%-20% of the elderly and approximately 16% of persons with chronic MTrPs 103,243.
B12 deficiencies can result in cognitive dysfunction,
degeneration of the spinal cord, and peripheral neuropathy, which is most likely linked to complaints of
diffuse myalgia seen in some patients. Serum levels of
vitamin B12 as high as 350 pg/ml may be associated with
a metabolic deficiency manifested by elevated serum or
urine methylmalonic acid or homocysteine and may be
clinically symptomatic244. However, there are patients
with normal levels of methylmalonic acid and homocysteine, who do present with metabolic abnormalities of
B12 function242. Folic acid is closely linked to vitamin
B12 and should be measured as well. While folic acid
is able to correct the pernicious anaemia associated
with vitamin B12 deficiency, it does not influence the
neuromuscular aspects. Iron deficiency in muscle occurs when ferritin is
depleted. Ferritin represents the tissue-bound non-essential iron stores in muscle, liver, and bone marrow
that supply the essential iron for oxygen transport and
iron-dependent enzymes. Iron is critical for the generation of energy through the cytochrome oxidase enzyme
system and a lack of iron may be a factor in the development and maintenance of MTrPs242. Interestingly, lowered
levels of cytochrome oxidase are common in patients
with myalgia140. Serum levels of 15-20 ng/ml indicate
a depletion of ferritin. Common symptoms are chronic
tiredness, coldness, extreme fatigue with exercise, and
muscle pain. Anaemia is common at levels of 10 ng/ml
or less. Although optimal levels of ferritin are unknown,
Gerwin242 suggested that levels below 50 ng/ml may be
clinically significant.
Close to 90% of patients with chronic musculoskeletal
pain may have vitamin D deficiency245. Vitamin D deficiencies are identified by measuring 25-OH vitamin D levels.
Levels above 20 ng/ml are considered normal, but Gerwin242
suggested that levels below 34 ng/ml may represent insufficiencies. Correction of insufficient levels of vitamin B12,
vitamin D, and iron levels may take many months, during
which patients may not see much improvement.
Myofascial Trigger Points: An Evidence-Informed Review / 213
Even when active MTrPs have been identified in a
particular patient, clinicians must always consider that
MTrPs may be secondary to metabolic insufficiencies
or other medical diagnoses. It is questionable whether
physical therapy and—as an integral part of physical
therapy management—manual therapy intervention
can be successful when patients have nutritional or
metabolic insufficiencies or deficiencies. A close working
relationship with physicians familiar with this body of
literature is essential. Therapists should consider the
possible interactions between arthrogenic or neurogenic
dysfunction and MTrPs4,5,118,133,246,247.
Clinically, physical therapists should address all
aspects of the dysfunction. There are many other conditions that feature muscle pain and MTrPs, including
hypothyroidism, systemic lupus erythematosis, Lyme
disease, babesiosis, ehrlichiosis, candida albicans infections, myoadenylate deaminase deficiency, hypoglycaemia,
and parasitic diseases such as fascioliasis, amoebiasis,
and giardia64, 242. Therapists should be familiar with the
symptoms associated with these medical diagnoses64.
Psychological stress may activate MTrPs. Electromyographic activity in MTrPs has been shown to increase
dramatically in response to mental and emotional stress,
whereas adjacent non-trigger point muscle EMG activity
remained normal 229, 230. Relaxation techniques, such as
autogenic relaxation, can diminish the electrical activity231. In addition, many patients with persistent MTrPs
are dealing with depression, anxiety, anger, and feelings
of hopelessness248. Pain-related fear and avoidance can
lead to the development and maintenance of chronic
pain249. Sleep disturbance can also be a major factor in
the perpetuation of musculoskeletal pain and must be
addressed. Sleep problems may be related to pain, apnea,
or to mood disorders like depression or anxiety. Management can be both pharmacologic and non-pharmacologic.
Pharmacologic treatment utilizes drugs that promote
normal sleep patterns and induce and maintain sleep
through the night without causing daytime sedation.
Non-pharmacologic treatment emphasizes sleep hygiene,
such as using the bed only for sleep and sex, and not
for reading, television viewing, and eating250. Therapists
must be sensitive to the impact of psychological and
emotional distress and refer patients to clinical social
workers or psychologists when appropriate.
The Role of Manual Therapy
Although the various management approaches are
beyond the scope of this article, manual therapy is one
of the basic treatment options and the role of orthopedic manual physical therapists cannot be overemphasized 82,158. Myofascial trigger points are treated with
manual techniques, spray and stretch, dry needling, or
injection therapy. Dry needling is within the scope of
physical therapy practice in many countries including
Canada, Spain, Ireland, South Africa, Australia, the
Netherlands, and Switzerland. In the United States,
the physical therapy boards of eight states have ruled
that physical therapists can engage in the practice
of dry needling: New Hampshire, Maryland, Virginia,
South Carolina, Georgia, Kentucky, New Mexico, and
Colorado80. A promising new development used in the
diagnosis and treatment of MTrPs involves shockwave
therapy, but as of yet, there are no controlled studies
substantiating its use251,252.
Summary
Although MTrPs are a common cause of pain and
dysfunction in persons with musculoskeletal injuries
and diagnoses, the importance of MTrPs is not obvious
from reviewing the orthopedic manual therapy literature. Current scientific evidence strongly supports that
awareness and a working knowledge of muscle dysfunction; in particular, MTrPs should be incorporated into
manual physical therapy practice consistent with the
IFOMT guidelines for clinical practice. While there are
still many unanswered questions with regard to explaining the etiology of MTrPs, this article provides manual
therapists with an up-to-date evidence-informed review
of the current scientific knowledge.
References
1.
2.
3.
4.
IFOMT. Available at: http://www.ifomt.org/ifomt/about/standards.
Accessed November 15, 2006.
Huijbregts PA. Muscle injury, regeneration, and repair. J Manual
Manipulative Ther 2001;9:9-16.
Urquhart DM, Hodges PW, Allen TJ, Story IH. Abdominal muscle
recruitment during a range of voluntary exercises. Man Ther
2005;10:144-153.
Fernández-de-las-Peñas C, Alonso-Blanco C, Alguacil-Diego IM,
Miangolarra-Page JC. Myofascial trigger points and postero-an-
214 / The Journal of Manual & Manipulative Therapy, 2006
5.
6.
terior joint hypomobility in the mid-cervical spine in subjects
presenting with mechanical neck pain: A pilot study. J Manual
Manipulative Ther 2006;14:88-94.
Fernández-de-las-Peñas C, Alonso-Blanco C, Miangolarra JC.
Myofascial trigger points in subjects presenting with mechanical
neck pain: A blinded, controlled study Man Ther (In press).
Lew PC, Lewis J, Story I. Inter-therapist reliability in locating
latent myofascial trigger points using palpation. Man Ther
1997;2:87-90.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
Fernández-de-las-Peñas C, Alonso-Blanco C, Cuadrado ML, Pareja
JA. Myofascial trigger points in the suboccipital muscles in
episodic tension-type headache. Man Ther 2006;11:225-230.
Moore A, Petty N. Evidence-based practice: Getting a grip and
finding a balance. Man Ther 2001;6:195-196.
Sackett DL, Rosenberg WM. The need for evidence-based medicine. J R Soc Med 1995;88:620-624.
Sackett DL, Rosenberg WM, Gray JA, Haynes RB, Richardson
WS. Evidence-based medicine: What it is and what it isn’t. BMJ
1996;312:71-72.
Pencheon D. What’s next for evidence-based medicine? EvidenceBased Healthcare Public Health 2005;9:319-321.
Cicerone KD. Evidence-based practice and the limits of rational
rehabilitation. Arch Phys Med Rehabil 2005;86:1073-1074.
Baldry PE. Acupuncture, Trigger Points and Musculoskeletal
Pain. Edinburgh, UK: Churchill Livingstone, 2005.
Simons DG. Muscle pain syndromes. Part 1. Am J Phys Med
1975;54:289-311.
Ruhmann W. The earliest book on rheumatism. Br J Rheumatism
1940;11:140-162.
Simons DG, Travell JG, Simons LS. Travell and Simons’ Myofascial Pain and Dysfunction: The Trigger Point Manual. Vol.
1. 2nd ed. Baltimore, MD: Williams & Wilkins, 1999.
Harden RN, Bruehl SP, Gass S, Niemiec C, Barbick B. Signs and
symptoms of the myofascial pain syndrome: A national survey
of pain management providers. Clin J Pain 2000;16:64-72.
Stockman R. The causes, pathology, and treatment of chronic
rheumatism. Edinburgh Med J 1904;15:107-116.
Strauss H. Űber die sogenannten‚ Rheumatische Muskelschwiele‘
[German; With regard to the so-called myogelosis]. Klin Wochenschr 1898;35:89-91,121-123.
Lange M. Die Muskelhärten (Myogelosen) [German; The Muscle
Hardenings (Myogeloses)]. Munich, Germany: J.F. Lehmann‘s
Verlag, 1931.
Travell J. Office Hours: Day and Night. The Autobiography of
Janet Travell, MD. New York, NY: World Publishing, 1968.
Kellgren JH. Deep pain sensibility. Lancet 1949;1:943-949.
Kellgren JH. Observations on referred pain arising from muscle.
Clin Sci 1938;3:175-190.
Kellgren JH. A preliminary account of referred pains arising
from muscle. British Med J 1938;1:325-327.
Simons DG. Cardiology and myofascial trigger points: Janet G.
Travell’s contribution. Tex Heart Inst J 2003;30(1):3-7.
Travell J. Basis for the multiple uses of local block of somatic
trigger areas (procaine infiltration and ethyl chloride spray).
Miss Valley Med 1949;71:13-22.
Travell J, Bobb AL. Mechanism of relief of pain in sprains by
local injection techniques. Fed Proc 1947;6:378.
Travell JG, Rinzler S, Herman M. Pain and disability of the
shoulder and arm: Treatment by intramuscular infiltration with
procaine hydrochloride. JAMA 1942;120:417-422.
Travell JG, Rinzler SH. The myofascial genesis of pain. Postgrad
Med 1952;11:452-434.
Chaitow L, DeLany J. Neuromuscular techniques in orthopedics.
Techniques in Orthopedics 2003;18(1):74-86.
Good MG. Five hundred cases of myalgia in the British army.
Ann Rheum Dis 1942;3:118-138.
Good MG. The role of skeletal muscle in the pathogenesis of
diseases. Acta Medica Scand 1950;138:285-292.
Gutstein M. Common rheumatism and physiotherapy. Br J Phys
Med 1940;3:46-50.
34. Gutstein M. Diagnosis and treatment of muscular rheumatism.
Br J Phys Med 1938;1:302-321.
35. Kelly M. The nature of fibrositis. I. The myalgic lesion and its
secondary effects: A reflex theory. Ann Rheum Dis 1945;5:1-7.
36. Kelly M. The nature of fibrositis. II. A study of the causation of
the myalgic lesion (rheumatic, traumatic, infective). Ann Rheum
Dis 1946;5:69-77.
37. Kelly M. The relief of facial pain by procaine (Novocaine) injections. J Am Geriatr Soc 1963;11:586-96.
38. Kelly M. The treatment of fibrositis and allied disorders by local
anesthesia. Med J Aust 1941;1:294-298.
39. Schneider M, Cohen J, Laws S. The Collected Writings of Nimmo
& Vannerson: Pioneers of Chiropractic Trigger Point Therapy.
Pittsburgh, PA: Schneider, 2001.
40. Cohen JH, Gibbons RW. Raymond L. Nimmo and the evolution
of trigger point therapy, 1929-1986. J Manipulative Physiol Ther
1998;21:167-172.
41. National Board of Chiropractic Examiners. Chiropractic Treatment Procedures. Greeley, CO: NBCE, 1993.
42. Mennell J. Spray-stretch for the relief of pain from muscle spasm
and myofascial trigger points. J Am Podiatry Assoc 1976;66:873876.
43. Mennell J. Myofascial trigger points as a cause of headaches. J
Manipulative Physiol Ther, 1989;12:308-313.
44. Travell W, Travell JG. Technic for reduction and ambulatory
treatment of sacroiliac displacement. Arch Phys Ther 1942;23:222232.
45. Paris SV. In the best interests of the patient. Phys Ther
2006;86:1541-1553.
46. Travell JG, Simons DG. Myofascial Pain and Dysfunction: The
Trigger Point Manual. Vol. 2. Baltimore, MD: Williams & Wilkins,
1992.
47. Travell JG, Simons DG. Myofascial Pain and Dysfunction: The
Trigger Point Manual. Vol. 1. Baltimore, MD: Williams & Wilkins,
1983.
48. Dejung B, Gröbli C, Colla F, Weissmann R. Triggerpunkttherapie
[German; Trigger Point Therapy]. Bern, Switzerland: Hans Huber,
2003.
49. Ferguson LW, Gerwin R. Clinical Mastery in the Treatment of
Myofascial Pain. Philadelphia, PA: Lippincott Williams & Wilkins,
2005.
50. Kostopoulos D, Rizopoulos K. The Manual of Trigger Point and
Myofascial Therapy. Thorofare, NJ: Slack, 2001.
51. Prateepavanich P. Myofascial Pain Syndrome: A Common Problem
in Clinical Practice. Bangkok, Thailand: Ammarind, 1999.
52. Rachlin ES, Rachlin IS. Myofascial Pain and Fibromyalgia:
Trigger Point Management. St. Louis, MO: Mosby, 2002.
53. Cardinal S. Points Détente et Acupuncture: Approche Neurophysiologique [French; Trigger Points and Acupuncture: Neurophysiological Approach]. Montreal, Canada: Centre Collégial
de Développement de Matériel Didactique, 2004.
54. Jonckheere PDM. Spieren en Dysfuncties, Trigger punten, Basisprincipes van de Myofasciale Therapie [Dutch; Muscles and
Dysfunctions, Basic Principles of Myofascial Therapy]. Brussels,
Belgium: Satas, 1993.
55. Lucas KR, Polus BI, Rich PS. Latent myofascial trigger points:
Their effect on muscle activation and movement efficiency. J
Bodywork Mov Ther 2004;8:160-166.
56. Weissmann RD. Überlegungen zur Biomechanik in der Myofaszialen Triggerpunkttherapie [German; Considerations with
regard to the biomechanics related to myofascial trigger point
Myofascial Trigger Points: An Evidence-Informed Review / 215
therapy]. Physiotherapie 2000;35(10):13-21.
57. Vecchiet L, Giamberardino MA, De Bigontina P. Comparative
sensory evaluation of parietal tissues in painful and nonpainful
areas in fibromyalgia and myofascial pain syndrome. In: Gebhart
GF, Hammond DL, Jensen TS, eds. Proceedings of the 7th World
Congress on Pain (Progress in Pain Research and Management).
Seattle, WA: IASP Press, 1994:177-185.
58. Vecchiet L, Giamberardino MA, Dragani L. Latent myofascial
trigger points: Changes in muscular and subcutaneous pain
thresholds at trigger point and target level. J Manual Medicine
1990;5:151-154.
59. Vecchiet L, Pizzigallo E, Iezzi S, Affaitati G, Vecchiet J, Giamberardino MA. Differentiation of sensitivity in different tissues and
its clinical significance. J Musculoskeletal Pain 1998;6:33-45.
60. Dommerholt J. Persistent myalgia following whiplash. Curr Pain
Headache Rep 2005;9:326-330.
61. Ge HY, Fernández-de-las-Peñas C, Arendt-Nielsen L. Sympathetic
facilitation of hyperalgesia evoked from myofascial tender and
trigger points in patients with unilateral shoulder pain. Clin
Neurophysiol 2006;117:1545-1550.
62. Lidbeck J. Central hyperexcitability in chronic musculoskeletal
pain: A conceptual breakthrough with multiple clinical implications. Pain Res Manag 2002;7(2):81-92.
63. Munglani R. Neurobiological mechanisms underlying chronic
whiplash associated pain: The peripheral maintenance of central
sensitization. J Musculoskeletal Pain 2000;8:169-178.
64. Dommerholt J, Issa T. Differential diagnosis: Myofascial pain.
In: Chaitow L, ed. Fibromyalgia Syndrome: A Practitioner’s
Guide to Treatment. Edinburgh, UK: Churchill Livingstone,
2003:149-177.
65. Gerwin RD, Shannon S, Hong CZ, Hubbard D, Gevirtz R. Interrater reliability in myofascial trigger point examination. Pain
1997;69(1-2):65-73.
66. Sciotti VM, Mittak VL, DiMarco L, Ford LM, Plezbert J, Santipadri
E, Wigglesworth J, Ball K. Clinical precision of myofascial trigger
point location in the trapezius muscle. Pain 2001;93(3):259266.
67. Franssen JLM. Handboek Oppervlakte Elektromyografie [Dutch;
Manual Surface Electromyography]. Utrecht, The Netherlands:
De Tijdstroom, 1995.
68. Janda V. Muscle spasm: A proposed procedure for differential
diagnosis. J Manual Med 1991;6:136-139.
69. Mense S. Pathophysiologic basis of muscle pain syndromes. In:
Fischer AA, ed. Myofascial Pain: Update in Diagnosis and Treatment. Philadelphia, PA: W.B. Saunders Company, 1997:23-53.
70. Headly BJ. Evaluation and treatment of myofascial pain syndrome
utilizing biofeedback. In: Cram JR, ed. Clinical Electromyography for Surface Recordings. Nevada City, NV: Clinical Resources,
1990:235-254.
71. Headly BJ. Chronic pain management. In: O’Sullivan SB, Schmitz
TS, eds. Physical Rehabilitation: Assessment and Treatment.
Philadelphia, PA: F.A. Davis Company, 1994:577-600.
72. Gerwin RD, Duranleau D. Ultrasound identification of the
myofascial trigger point. Muscle Nerve 1997;20:767-768.
73. Hong CZ. Persistence of local twitch response with loss of
conduction to and from the spinal cord. Arch Phys Med Rehabil
1994;75:12-16.
74. Hong C-Z, Torigoe Y. Electrophysiological characteristics of
localized twitch responses in responsive taut bands of rabbit
skeletal muscle. J Musculoskeletal Pain 1994;2:17-43.
75. Hong C-Z, Yu J. Spontaneous electrical activity of rabbit trig-
216 / The Journal of Manual & Manipulative Therapy, 2006
76.
77.
78.
79.
80.
81.
82.
83.
84.
85.
86.
87.
88.
89.
90.
91.
92.
93.
94.
ger spot after transection of spinal cord and peripheral nerve.
J Musculoskeletal Pain 1998;6(4):45-58.
Fricton JR, Auvinen MD, Dykstra D, Schiffman E. Myofascial
pain syndrome: Electromyographic changes associated with local
twitch response. Arch Phys Med Rehabil 1985;66:314-317.
Simons DG, Dexter JR. Comparison of local twitch responses
elicited by palpation and needling of myofascial trigger points.
J Musculoskeletal Pain 1995;3:49-61.
Wang F, Audette J. Electrophysiological characteristics of the local
twitch response with active myofascial pain of neck compared
with a control group with latent trigger points. Am J Phys Med
Rehabil 2000;79:203.
Audette JF, Wang F, Smith H. Bilateral activation of motor unit
potentials with unilateral needle stimulation of active myofascial
trigger points. Am J Phys Med Rehabil 2004;83:368-374, quiz
375-377,389.
Dommerholt J. Dry needling in orthopedic physical therapy
practice. Orthop Phys Ther Pract 2004;16(3):15-20.
Hong C-Z. Lidocaine injection versus dry needling to myofascial
trigger point: The importance of the local twitch response. Am
J Phys Med Rehabil 1994;73:256-263.
Gerwin RD, Dommerholt J. Treatment of myofascial pain syndromes. In: Boswell MV, Cole BE, eds. Weiner’s Pain Management: A Practical Guide for Clinicians. Boca Raton, FL: CRC
Press, 2006:477-492.
Vecchiet L, Dragani L, De Bigontina P, Obletter G, Giamberardino
MA. Experimental referred pain and hyperalgesia from muscles
in humans. In: Vecchiet L, et al, eds. New Trends in Referred
Pain and Hyperalgesia. Amsterdam, The Netherlands: Elsevier
Science, 1993:239-249.
Vecchiet L, Giamberardino MA. Referred pain: Clinical significance,
pathophysiology and treatment. In: Fischer AA, ed. Myofascial
Pain: Update in Diagnosis and Treatment. Philadelphia, PA:
W.B. Saunders Company, 1997:119-136.
Mense S. The pathogenesis of muscle pain. Curr Pain Headache
Rep 2003;7:419-425.
Mense S. Referral of muscle pain: New aspects. Amer Pain Soc
J 1994;3:1-9.
Fernández-de-las-Peñas CF, Cuadrado ML, Gerwin RD, Pareja
JA. Referred pain from the trochlear region in tension-type
headache: A myofascial trigger point from the superior oblique
muscle. Headache 2005;45:731-737.
Fernández-de-las-Peñas C, Cuadrado ML, Gerwin RD, Pareja
JA. Myofascial disorders in the trochlear region in unilateral
migraine: A possible initiating or perpetuating factor. Clin J
Pain 2006;22:548-553.
Hwang M, Kang YK, Kim DH. Referred pain pattern of the
pronator quadratus muscle. Pain 2005;116:238-242.
Hwang M, Kang YK, Shin JY, Kim DH. Referred pain pattern of
the abductor pollicis longus muscle. Am J Phys Med Rehabil
2005;84:593-597.
Hong CZ, Kuan TS, Chen JT, Chen SM. Referred pain elicited
by palpation and by needling of myofascial trigger points: A
comparison. Arch Phys Med Rehabil 1997;78:957-960.
Dwyer A, Aprill C, Bogduk N. Cervical zygapophyseal joint pain
patterns. I: A study in normal volunteers. Spine 1990;15:453457.
Giamberardino MA, Vecchiet L. Visceral pain, referred hyperalgesia and outcome: New concepts. Eur J Anaesthesiol (Suppl)
1995;10:61-66.
Scudds RA, Landry M, Birmingham T, Buchan J, Griffin K. The
frequency of referred signs from muscle pressure in normal
healthy subjects (abstract). J Musculoskeletal Pain 1995;3 (Suppl.
1):99.
95. Torebjörk HE, Ochoa JL, Schady W. Referred pain from intraneural stimulation of muscle fascicles in the median nerve. Pain
1984;18:145-156.
96. Gibson W, Arendt-Nielsen L, Graven-Nielsen T. Referred pain
and hyperalgesia in human tendon and muscle belly tissue.
Pain 2006;120(1-2):113-123.
97. Gibson W, Arendt-Nielsen L, Graven-Nielsen T. Delayed onset
muscle soreness at tendon-bone junction and muscle tissue is
associated with facilitated referred pain. Exp Brain Res 2006
(In press).
98. Hendler NH, Kozikowski JG. Overlooked physical diagnoses in
chronic pain patients involved in litigation. Psychosomatics
1993;34:494-501.
99. Weiner DK, Sakamoto S, Perera S, Breuer P. Chronic low back
pain in older adults: Prevalence, reliability, and validity of physical examination findings. J Am Geriatr Soc 2006;54:11-20.
100. Calandre EP, Hidalgo J, Garcia-Leiva JM, Rico-Villademoros F.
Trigger point evaluation in migraine patients: An indication of
peripheral sensitization linked to migraine predisposition? Eur
J Neurol 2006;13:244-249.
101. Headache Classification Subcommittee of the International
Headache Society: The international classification of headache
disorders. Cephalalgia 2004;24(Suppl 1):9-160.
102. Fricton JR, Kroening R, Haley D, Siegert R. Myofascial pain syndrome of the head and neck: A review of clinical characteristics
of 164 patients. Oral Surg Oral Med Oral Pathol 1985;60:615623.
103. Gerwin R. A study of 96 subjects examined both for fibromyalgia and myofascial pain (abstract). J Musculoskeletal Pain
1995;3(Suppl 1):121.
104. Rosomoff HL, Fishbain DA, Goldberg N, Rosomoff RS. Myofascial
findings with patients with chronic intractable benign pain of
the back and neck. Pain Management 1989;3:114-118.
105. Skootsky SA, Jaeger B, Oye RK. Prevalence of myofascial pain
in general internal medicine practice. West J Med 1989;151:157160.
106. Chaiamnuay P, Darmawan J, Muirden KD, Assawatanabodee P.
Epidemiology of rheumatic disease in rural Thailand: A WHOILAR COPCORD study. Community Oriented Programme for
the Control of Rheumatic Disease. J Rheumatol 1998;25:13821387.
107. Graff-Radford B. Myofascial trigger points: Their importance and
diagnosis in the dental office. J Dent Assoc S Afr 1984;39:249253.
108. Bajaj P, Bajaj P, Graven-Nielsen T, Arendt-Nielsen L. Trigger
points in patients with lower limb osteoarthritis. J Musculoskeletal Pain 2001;9(3):17-33.
109. Hsueh TC, Yu S, Kuan TS, Hong C-Z. Association of active
myofascial trigger points and cervical disc lesions. J Formosa
Med Assoc 1998;97(3):174-80.
110. Wang C-F, Chen M, Lin M-T, Kuan T-S, Hong C-Z. Teres minor
tendinitis manifested with chronic myofascial pain syndrome
in the scapular muscles: A case report. J Musculoskeletal Pain
2006;14(1):39-43.
111. Fricton JR. Etiology and management of masticatory myofascial
pain. J Musculoskeletal Pain 1999;7(1/2):143-160.
112. Teachey WS. Otolaryngic myofascial pain syndromes. Curr Pain
Headache Rep 2004;8:457-462.
113. Dommerholt J. El sindrome de dolor miofascial en la region
craneomandibular. [Spanish; Myofascial pain syndrome in
the craniomandibular region]. In: Padrós Serrat E, ed. Bases
diagnosticas, terapeuticas y posturales del functionalismo
craniofacial. Madrid, Spain: Ripano, 2006:564-581.
114. Hesse J, Mogelvang B, Simonsen H. Acupuncture versus metoprolol in migraine prophylaxis: A randomized trial of trigger
point inactivation. J Intern Med 1994;235:451-456.
115. Skubick DL, Clasby R, Donaldson CC, Marshall WM. Carpal
tunnel syndrome as an expression of muscular dysfunction in
the neck. J Occupational Rehab 1993;3:31-43.
116. Treaster D, Marras WS, Burr D, Sheedy JE, Hart D. Myofascial
trigger point development from visual and postural stressors
during computer work. J Electromyogr Kinesiol 2006;16:115124.
117. Baker BA. The muscle trigger: Evidence of overload injury. J
Neurol Orthop Med Surg 1986;7:35-44.
118. Fruth SJ. Differential diagnosis and treatment in a patient with
posterior upper thoracic pain. Phys Ther 2006;86:254-268.
119. Doggweiler-Wiygul R. Urologic myofascial pain syndromes. Curr
Pain Headache Rep 2004;8:445-451.
120. Jarrell J. Myofascial dysfunction in the pelvis. Curr Pain Headache Rep 2004;8:452-456.
121. Jarrell JF, Vilos GA, Allaire C, Burgess S, Fortin C, Gerwin R,
Lapensee L, Lea RH, Leyland NA, Martyn P, Shenassa H, Taenzer
P, Abu-Rafea B. Consensus guidelines for the management of
chronic pelvic pain. J Obstet Gynaecol Can 2005;27:869-887.
122. Weiss JM. Pelvic floor myofascial trigger points: Manual therapy
for interstitial cystitis and the urgency-frequency syndrome. J
Urol 2001;166:2226-2231.
123. Dommerholt J. Muscle pain syndromes. In: Cantu RI, Grodin
AJ, eds. Myofascial Manipulation. Gaithersburg, MD: Aspen,
2001:93-140.
124. Weiner DK, Schmader KE. Postherpetic pain: More than sensory
neuralgia? Pain Med 2006;7:243-249; discussion 250.
125. Chen SM, Chen JT, Kuan TS, Hong C-Z. Myofascial trigger points
in intercostal muscles secondary to herpes zoster infection of the
intercostal nerve. Arch Phys Med Rehabil 1998;79:336-338.
126. Dommerholt J. Complex regional pain syndrome. Part 1: History, diagnostic criteria and etiology. J Bodywork Mov Ther
2004;8:167-177.
127. Rashiq S, Galer BS. Proximal myofascial dysfunction in complex
regional pain syndrome: A retrospective prevalence study. Clin
J Pain,1999;15:151-153.
128. Prateepavanich P, Kupniratsaikul VC. The relationship between
myofascial trigger points of gastrocnemius muscle and nocturnal
calf cramps. J Med Assoc Thailand 1999;82:451-459.
129. Kern KU, Martin C, Scheicher S, Muller H. Auslosung von Phantomschmerzen und -sensationen durch Muskuläre Stumpftriggerpunkte nach Beinamputationen [German; Referred pain
from amputation stump trigger points into the phantom limb].
Schmerz 2006;20:300-306.
130. Kern U, Martin C, Scheicher S, Műller H. Does botulinum toxin
A make prosthesis use easier for amputees? J Rehabil Med
2004;36:238-239.
131. Longbottom J. A case report of postulated “Barré Liéou syndrome.” Acupunct Med 2005;23:34-38.
132. Stellon A. Neurogenic pruritus: An unrecognized problem? A
retrospective case series of treatment by acupuncture. Acupunct
Med 2002;20:186-190.
133. Fernández-de-las-Peñas C, Fernández-Carnero J, Miangolarra-
Myofascial Trigger Points: An Evidence-Informed Review / 217
Page JC. Musculoskeletal disorders in mechanical neck pain:
Myofascial trigger points versus cervical joint dysfunction. J
Musculoskeletal Pain 2005;13(1):27-35.
134. Hägg GM. Ny förklaringsmodell för muskelskador vid statisk
belastning i skuldra och nacke [Swedish; New explanation
for muscle damage as a result of static loads in the neck and
shoulder]. Arbete Människa Miljö 1988;4:260-262.
135. Henneman E, Somjen G, Carpenter DO. Excitability and inhibitability of motoneurons of different sizes. J Neurophysiol
1965;28:599-620.
136. Hägg GM. The Cinderella Hypothesis. In: Johansson H, et al, eds.
Chronic Work-Related Myalgia. Gävle, Sweden: Gävle University
Press, 2003:127-132.
137. Armstrong RB. Initial events in exercise-induced muscular
injury. Med Sci Sports Exerc 1990;22:429-435.
138. Hägg GM. Human muscle fibre abnormalities related to occupational load. Eur J Appl Physiol 2000;83(2-3):159-165.
139. Kadi F, Hagg G, Hakansson R, Holmner S, Butler-Browne GS,
Thornell LE. Structural changes in male trapezius muscle with
work-related myalgia. Acta Neuropathol (Berl) 1998;95:352360.
140. Kadi F, Waling K, Ahlgren C, Sundelin G, Holmner S, ButlerBrowne GS, Thornell LE. Pathological mechanisms implicated
in localized female trapezius myalgia. Pain 1998;78:191-196.
141. Larsson B, Bjork J, Kadi F, Lindman R, Gerdle B. Blood supply
and oxidative metabolism in muscle biopsies of female cleaners
with and without myalgia. Clin J Pain 2004;20:440-446.
142. Henriksson KG, Bengtsson A, Lindman R, Thornell LE. Morphological changes in muscle in fibromyalgia and chronic shoulder
myalgia. In: Værøy H, Merskey H, eds. Progress in Fibromyalgia
and Myofascial Pain. Amsterdam, The Netherlands: Elsevier,
1993:61-73.
143. Lexell J, Jarvis J, Downham D, Salmons S. Stimulation-induced
damage in rabbit fast-twitch skeletal muscles: A quantitative
morphological study of the influence of pattern and frequency.
Cell Tissue Res 1993;273:357-362.
144. Gissel H. Ca2+ accumulation and cell damage in skeletal muscle
during low frequency stimulation. Eur J Appl Physiol 2000;83(23):175-180.
145. Gissel H, Clausen T. Excitation-induced Ca(2+) influx in rat
soleus and EDL muscle: Mechanisms and effects on cellular
integrity. Am J Physiol Regul Integr Comp Physiol 2000;279:
R917-924.
146. Febbraio MA, Pedersen BK. Contraction-induced myokine
production and release: Is skeletal muscle an endocrine organ?
Exerc Sport Sci Rev 2005;33(3):114-119.
147. Pedersen BK, Febbraio M. Muscle-derived interleukin-6: A possible link between skeletal muscle, adipose tissue, liver, and
brain. Brain Behav Immun 2005;19:371-376.
148. Forsman M, Birch L, Zhang Q, Kadefors R. Motor unit recruitment in the trapezius muscle with special reference to coarse
arm movements. J Electromyogr Kinesiol 2001;11:207-216.
149. Forsman M, Kadefors R, Zhang Q, Birch L, Palmerud G. Motorunit recruitment in the trapezius muscle during arm movements
and in VDU precision work. Int J Ind Ergon 1999;24:619-630.
150. Forsman M, Taoda K, Thorn S, Zhang Q. Motor-unit recruitment
during long-term isometric and wrist motion contractions:
A study concerning muscular pain development in computer
operators. Int J Ind Ergon 2002;30:237-250.
151. Zennaro D, Laubli T, Krebs D, Klipstein A, Krueger H. Continuous, intermitted and sporadic motor unit activity in the trape-
218 / The Journal of Manual & Manipulative Therapy, 2006
zius muscle during prolonged computer work. J Electromyogr
Kinesiol 2003;13:113-124.
152. Zennaro D, Laubli T, Krebs D, Krueger H, Klipstein A. Trapezius
muscle motor unit activity in symptomatic participants during
finger tapping using properly and improperly adjusted desks.
Hum Factors 2004;46:252-66.
153. Andersen JH, Kærgaard A, Rasmussen K. Myofascial pain in
different occupational groups with monotonous repetitive work
(abstract). J Musculoskeletal Pain 1995;3(Suppl 1):57.
154. Chen S-M, Chen J-T, Kuan T-S, Hong J, Hong C-Z. Decrease in
pressure pain thresholds of latent myofascial trigger points in
the middle finger extensors immediately after continuous piano
practice. J Musculoskeletal Pain 2000;8(3):83-92.
155. Otten E. Concepts and models of functional architecture in
skeletal muscle. Exerc Sport Sci Rev 1988;16:89-137.
156. Sjogaard G, Lundberg U, Kadefors R. The role of muscle activity
and mental load in the development of pain and degenerative
processes at the muscle cell level during computer work. Eur
J Appl Physiol 2000;83(2-3):99-105.
157. Sjogaard G, Sogaard K. Muscle injury in repetitive motion
disorders. Clin Orthop 1998;351:21-31.
158. Simons DG. Understanding effective treatments of myofascial
trigger points. J Bodywork Mov Ther 2002;6:81-88.
159. Proske U, Morgan DL. Stiffness of cat soleus muscle and tendon
during activation of part of muscle. J Neurophysiol 1984;52:459468.
160. Trotter JA. Functional morphology of force transmission in
skeletal muscle: A brief review. Acta Anat (Basel) 1993;146(4):205222.
161. Monti RJ, Roy RR, Hodgson JA, Edgerton VR. Transmission
of forces within mammalian skeletal muscles. J Biomech
1999;32:371-380.
162. Bodine SC, Roy RR, Eldred E, Edgerton VR. Maximal force as a
function of anatomical features of motor units in the cat tibialis
anterior. J Neurophysiol 1987;57:1730-1745.
163. Ounjian M, Roy RR, Eldred E, Garfinkel A, Payne JR, Armstrong
A, Toga AW, Edgerton VR. Physiological and developmental
implications of motor unit anatomy. J Neurobiol 1991;22:547559.
164. Petit J, Filippi GM, Emonet-Denand F, Hunt CC, Laporte Y.
Changes in muscle stiffness produced by motor units of different types in peroneus longus muscle of cat. J Neurophysiol
1990;63:190-197.
165. Petit J, Filippi GM, Gioux M, Hunt CC, Laporte Y. Effects of
tetanic contraction of motor units of similar type on the initial
stiffness to ramp stretch of the cat peroneus longus muscle. J
Neurophysiol 1990;64:1724-1732.
166. Altringham JD, Bottinelli R. The descending limb of the sarcomere length-force relation in single muscle fibres of the frog.
J Muscle Res Cell Motil 1985;6:585-600.
167. Street SF. Lateral transmission of tension in frog myofibers: A
myofibrillar network and transverse cytoskeletal connections
are possible transmitters. J Cell Physiol 1983;114:346-364.
168. Denoth J, Stussi E, Csucs G, Danuser G. Single muscle fiber
contraction is dictated by inter-sarcomere dynamics. J Theor
Biol 2002;216(1):101-122.
169. Dommerholt J, Royson MW, Whyte-Ferguson L. Neck pain and
dysfunction following whiplash. In: Whyte-Ferguson L, Gerwin
RD, eds. Clinical Mastery of Myofascial Pain Syndrome. Baltimore, MD: Lippincott, Williams & Wilkins, 2005:57-89.
170. Jull GA. Deep cervical flexor muscle dysfunction in whiplash. J
Musculoskeletal Pain 2000;8(1/2):143-154.
171. Kumar S, Narayan Y, Amell T. Analysis of low velocity frontal
impacts. Clin Biomech 2003;18:694-703.
172. Sterling M, Jull G, Vicenzino B, Kenardy J, Darnell R. Development of motor system dysfunction following whiplash injury.
Pain 2003;103(1-2):65-73.
173. Sterling M, Jull G, Vicenzino B, Kenardy J, Darnell R. Physical
and psychological factors predict outcome following whiplash
injury. Pain 2005;114(1-2):141-148.
174. Schuller E, Eisenmenger W, Beier G. Whiplash injury in low speed
car accidents. J Musculoskeletal Pain 2000;8(1/2):55-67.
175. Simons DG. Triggerpunkte und Myogelose [German; Trigger
points and myogeloses]. Manuelle Medizin 1997;35:290-294.
176. Gerwin RD, Dommerholt J. Myofascial trigger points in chronic
cervical whiplash syndrome. J Musculoskeletal Pain 1998;6(Suppl.
2):28.
177. Fernández-de-las-Peñas C, Fernández-Carnero J, Palomequedel-Cerro L, Miangolarra-Page JC. Manipulative treatment vs.
conventional physiotherapy treatment in whiplash injury: A
randomized controlled trial. J Whiplash Rel Disord 2004;3(2):7390.
178. Fernández-de-las-Peñas C, Palomeque-del-Cerro L, FernándezCarnero J. Manual treatment of post-whiplash injury. J Bodywork
Mov Ther 2005;9:109-119.
179. Simons DG, Travell J, Myofascial trigger points: A possible
explanation. Pain 1981;10:106-109.
180. Sterling M, Jull G, Vicenzino B, Kenardy J. Sensory hypersensitivity occurs soon after whiplash injury and is associated with
poor recovery. Pain 2003;104:509-517.
181. Scott D, Jull G, Sterling M. Widespread sensory hypersensitivity is a feature of chronic whiplash-associated disorder but not
chronic idiopathic neck pain. Clin J Pain 2005;21:175-181.
182. Vecchiet L, Vecchiet J, Bellomo R, Giamberardino MA. Muscle pain
from physical exercise. J Musculoskeletal Pain 1999;7(1/2):4353.
183. Simons DG. Review of enigmatic MTrPs as a common cause of
enigmatic musculoskeletal pain and dysfunction. J Electromyogr
Kinesiol 2004;14:95-107.
184. Gerwin RD, Dommerholt J, Shah J. An expansion of Simons’
integrated hypothesis of trigger point formation. Curr Pain
Headache Rep 2004;8:468-475.
185. Newham DJ, Jones DA, Clarkson PM. Repeated high-force eccentric exercise: Effects on muscle pain and damage. J Appl
Physiol 1987;63:1381-1386.
186. Fridén J, Lieber RL. Segmental muscle fiber lesions after repetitive eccentric contractions. Cell Tissue Res 1998;293:165-171.
187. Stauber WT, Clarkson PM, Fritz VK, Evans WJ. Extracellular
matrix disruption and pain after eccentric muscle action. J Appl
Physiol 1990;69:868-874.
188. Graven-Nielsen T, Arendt-Nielsen L. Induction and assessment
of muscle pain, referred pain, and muscular hyperalgesia. Curr
Pain Headache Rep 2003;7:443-451.
189. Lieber RL, Shah S, Fridén J. Cytoskeletal disruption after eccentric contraction-induced muscle injury. Clin Orthop 2002;403:
S90-S99.
190. Lieber RL, Thornell LE, Fridén J. Muscle cytoskeletal disruption
occurs within the first 15 min of cyclic eccentric contraction.
J Appl Physiol 1996;80:278-284.
191. Barash IA, Peters D, Fridén J, Lutz GJ, Lieber RL. Desmin
cytoskeletal modifications after a bout of eccentric exercise
in the rat. Am J Physiol Regul Integr Comp Physiol 2002;283:
R958-R963.
192. Thompson JL, Balog EM, Fitts RH, Riley DA. Five myofibrillar
lesion types in eccentrically challenged, unloaded rat adductor
longus muscle: A test model. Anat Rec 1999;254:39-52.
193. Lieber RL, Fridén J. Mechanisms of muscle injury gleaned from
animal models. Am J Phys Med Rehabil 2002;81(11 Suppl):S70S79.
194. Peters D, Barash IA, Burdi M, Yuan PS, Mathew L, Fridén J,
Lieber RL. Asynchronous functional, cellular and transcriptional
changes after a bout of eccentric exercise in the rat. J Physiol
2003;553(Pt 3):947-957.
195. Bowers EJ, Morgan DL, Proske U. Damage to the human quadriceps muscle from eccentric exercise and the training effect.
J Sports Sci 2004;22(11-12):1005-1014.
196. Byrne C, Twist C, Eston R. Neuromuscular function after exerciseinduced muscle damage: Theoretical and applied implications.
Sports Med 2004;34(1):49-69.
197. Hamlin MJ, Quigley BM. Quadriceps concentric and eccentric
exercise. 2: Differences in muscle strength, fatigue and EMG
activity in eccentrically-exercised sore and non-sore muscles.
J Sci Med Sport 2001;4(1):104-115.
198. Pearce AJ, Sacco P, Byrnes ML, Thickbroom GW, Mastaglia FL.
The effects of eccentric exercise on neuromuscular function of
the biceps brachii. J Sci Med Sport 1998;1(4):236-244.
199. Shah JP, Phillips TM, Danoff JV, Gerber LH. An in-vivo microanalytical technique for measuring the local biochemical milieu
of human skeletal muscle. J Appl Physiol 2005;99:1980-1987.
200. Itoh K, Okada K, Kawakita K. A proposed experimental model of
myofascial trigger points in human muscle after slow eccentric
exercise. Acupunct Med 2004;22(1):2-12; discussion 12-13.
201. Brückle W, Sückfull M, Fleckenstein W, Weiss C, Müller W.
Gewebe-pO2-Messung in der verspannten Rückenmuskulatur
(m. erector spinae) [German; Tissue pO2 in hypertonic back
muscles]. Z. Rheumatol 1990;49:208-216.
202. McCleskey EW, Gold MS. Ion channels of nociception. Annu Rev
Physiol 1999;61:835-856.
203. Sluka KA, Kalra A, Moore SA. Unilateral intramuscular injections
of acidic saline produce a bilateral, long-lasting hyperalgesia.
Muscle Nerve 2001;24:37-46.
204. Sluka KA, Price MP, Breese NM, Stucky CL, Wemmie JA, Welsh
MJ. Chronic hyperalgesia induced by repeated acid injections in
muscle is abolished by the loss of ASIC3, but not ASIC1. Pain
2003;106:229-239.
205. Weeks VD. Travell J. How to give painless injections. In: AMA
Scientific Exhibits. New York, NY: Grune & Stratton, 1957:318322.
206. Hubbard DR, Berkoff GM. Myofascial trigger points show spontaneous needle EMG activity. Spine 1993;18:1803-1807.
207. Couppé C, Midttun A, Hilden J, Jørgensen U, Oxholm P, FuglsangFrederiksen A. Spontaneous needle electromyographic activity in
myofascial trigger points in the infraspinatus muscle: A blinded
assessment. J Musculoskeletal Pain 2001;9(3):7-17.
208. Simons DG. Do endplate noise and spikes arise from normal
motor endplates? Am J Phys Med Rehabil 2001;80:134-140.
209. Simons DG, Hong C-Z, Simons LS. Endplate potentials are
common to midfiber myofascial trigger points. Am J Phys Med
Rehabil 2002;81:212-222.
210. Macgregor J, Graf von Schweinitz D. Needle electromyographic
activity of myofascial trigger points and control sites in equine
cleidobrachialis muscle: An observational study. Acupunct Med
2006;24(2):61-70.
Myofascial Trigger Points: An Evidence-Informed Review / 219
211. Mense S. Neurobiological basis for the use of botulinum toxin
in pain therapy. J Neurol 2004;251(Suppl. 1):I1-7.
212. De Paiva A, Meunier FA, Molgo J, Aoki KR, Dolly JO. Functional
repair of motor endplates after botulinum neurotoxin type A
poisoning: Biphasic switch of synaptic activity between nerve
sprouts and their parent terminals. Proc Natl Acad Sci USA
1999;96:3200-3205.
213. Chen BM, Grinnell AD. Kinetics, Ca2+ dependence, and biophysical properties of integrin-mediated mechanical modulation of
transmitter release from frog motor nerve terminals. J Neurosci
1997;17:904-916.
214. Grinnell AD, Chen BM, Kashani A, Lin J, Suzuki K, Kidokoro
Y. The role of integrins in the modulation of neurotransmitter
release from motor nerve terminals by stretch and hypertonicity.
J Neurocytol 2003;32(5-8):489-503.
215. Kashani AH, Chen BM, Grinnell AD. Hypertonic enhancement
of transmitter release from frog motor nerve terminals: Ca2+
independence and role of integrins. J Physiol 2001;530(Pt
2):243-252.
216. Wang K, Yu L. Emerging concepts of muscle contraction and
clinical implications for myofascial pain syndrome (abstract). In:
Focus on Pain. Mesa, AZ: Janet G. Travell, MD Seminar Series,
2000.
217. Bukharaeva EA, Salakhutdinov RI, Vyskocil F, Nikolsky EE.
Spontaneous quantal and non-quantal release of acetylcholine at mouse endplate during onset of hypoxia. Physiol Res
2005;54:251-255.
218. Hoheisel U, Mense S, Simons D, Yu X-M. Appearance of new
receptive fields in rat dorsal horn neurons following noxious
stimulation of skeletal muscle: A model for referral of muscle
pain? Neurosci Lett 1993;153:9-12.
219. Simons DG, Stolov WC. Microscopic features and transient
contraction of palpable bands in canine muscle. Am J Phys Med
1976;55:65-88.
220. Reitinger A, Radner H, Tilscher H, Hanna M, Windisch A, Feigl
W. Morphologische Untersuchung an Triggerpunkten [German;
Morphological investigation of trigger points]. Manuelle Medizin
1996;34:256-262.
221. Windisch A, Reitinger A, Traxler H, Radner H, Neumayer C, Feigl
W, Firbas W. Morphology and histochemistry of myogelosis. Clin
Anat 1999;12:266-271.
222. Mense S, Simons DG, Hoheisel U, Quenzer B. Lesions of rat
skeletal muscle after local block of acetylcholinesterase and neuromuscular stimulation. J Appl Physiol 2003;94:2494-2501.
223. Pongratz D. Neuere Ergebnisse zur Pathogenese Myofaszialer Schmerzsyndrom [German; New findings with regard to
the etiology of myofascial pain syndrome]. Nervenheilkunde
2002;21(1):35-37.
224. Pongratz DE, Späth M. Morphologic aspects of muscle pain
syndromes. In: Fischer AA, ed. Myofascial Pain: Update in
Diagnosis and Treatment. Philadelphia, PA: W.B. Saunders
Company, 1997:55-68.
225. Gariphianova MB. The ultrastructure of myogenic trigger points
in patients with contracture of mimetic muscles (abstract). J
Musculoskeletal Pain 1995;3(Suppl 1):23.
226. Glogowsky C, Wallraff J. Ein Beitrag zur Klinik und Histologie der
Muskelhärten (Myogelosen) [German; A contribution on clinical
aspects and histology of myogeloses]. Z Orthop 1951;80:237268.
227. Fassbender HG. Morphologie und Pathogenese des Weichteilrheumatismus [German; Morphology and etiology of soft
220 / The Journal of Manual & Manipulative Therapy, 2006
tissue rheumatism]. Z Rheumaforsch 1973;32:355-374.
228. Hubbard DR. Chronic and recurrent muscle pain: Pathophysiology and treatment, and review of pharmacologic studies. J
Musculoskeletal Pain 1996;4:123-143.
229. Lewis C, Gevirtz R, Hubbard D, Berkoff G. Needle trigger point
and surface frontal EMG measurements of psychophysiological
responses in tension-type headache patients. Biofeedback &
Self-Regulation 1994;3:274-275.
230. McNulty WH, Gevirtz RN, Hubbard DR, Berkoff GM. Needle
electromyographic evaluation of trigger point response to a
psychological stressor. Psychophysiology 1994;31:313-316.
231. Banks SL, Jacobs DW, Gevirtz R, Hubbard DR. Effects of autogenic relaxation training on electromyographic activity in active
myofascial trigger points. J Musculoskeletal Pain 1998;6(4):2332.
232. Chen JT, Chen SM, Kuan TS, Chung KC, Hong C-Z. Phentolamine
effect on the spontaneous electrical activity of active loci in a
myofascial trigger spot of rabbit skeletal muscle. Arch Phys Med
Rehabil 1998;79:790-794.
233. Chen SM, Chen JT, Kuan TS, Hong C-Z. Effect of neuromuscular
blocking agent on the spontaneous activity of active loci in a
myofascial trigger spot of rabbit skeletal muscle. J Musculoskeletal Pain 1998;6(Suppl. 2):25.
234. Bowman WC, Marshall IG, Gibb AJ, Harborne AJ. Feedback
control of transmitter release at the neuromuscular junction.
Trends Pharmacol Sci 1988;9(1):16-20.
235. Cunha FQ, Lorenzetti BB, Poole S, Ferreira SH. Interleukin-8 as
a mediator of sympathetic pain. Br J Pharmacol 1991;104:765767.
236. Verri WA, Jr., Cunha TM, Parada CA, Poole S, Cunha FQ, Ferreira SH. Hypernociceptive role of cytokines and chemokines:
Targets for analgesic drug development? Pharmacol Ther 2006
(In press).
237. Poole S, de Queiroz Cunha F, Ferreira SH. Hyperalgesia from
subcutaneous cytokines. In: Watkins LR, Maier SF, eds. Cytokines
and Pain. Basel, Switzerland: Birkhaueser, 1999:59-87.
238. Fernandez HL, Hodges-Savola CA. Physiological regulation of
G4 AChE in fast-twitch muscle: Effects of exercise and CGRP.
J Appl Physiol 1996;80:357-362.
239. Hodges-Savola CA, Fernandez HL. A role for calcitonin generelated peptide in the regulation of rat skeletal muscle G4
acetylcholinesterase. Neurosci Lett 1995;190(2):117-20.
240. Fernandez-de-las-Penas C, Alonso-Blanco C, Cuadrado ML,
Gerwin RD, Pareja JA. Trigger points in the suboccipital muscles
and forward head posture in tension-type headache. Headache
2006;46:454-460.
241. Gerwin RD. Factores que promueven la persistencia de mialgia
en el síndrome de dolor miofascial y en la fibromyalgia [Spanish; Factors that promote the continued existence of myalgia
in myofascial pain syndrome and fibromyalgia]. Fisioterapia
2005;27(2):76-86.
242. Gerwin RD. A review of myofascial pain and fibromyalgia: Factors
that promote their persistence. Acupunct Med 2005;23(3):121134.
243. Andres E, Loukili NH, Noel E, Kaltenbach G, Abdelgheni MB,
Perrin AE, Noblet-Dick M, Maloisel F, Schlienger JL, Blickle JF.
Vitamin B12 (cobalamin) deficiency in elderly patients. CMAJ
2004;171:251-259.
244. Pruthi RK, Tefferi A. Pernicious anemia revisited. Mayo Clin
Proc 1994;69:144-150.
245. Plotnikoff GA, Quigley JM. Prevalence of severe hypovitaminosis
D in patients with persistent, nonspecific musculoskeletal pain.
Mayo Clin Proc 2003;78:1463-1470.
246. Bogduk N, Simons DG. Neck pain: Joint pain or trigger points.
In: Værøy H, Merskey H, eds. Progress in Fibromyalgia and
Myofascial Pain. Amsterdam, The Netherlands: Elsevier, 1993:267273.
247. Padamsee M, Mehta N, White GE. Trigger point injection: A
neglected modality in the treatment of TMJ dysfunction. J Pedod
1987;12(1):72-92.
248. Linton SJ. A review of psychological risk factors in back and
neck pain. Spine 2000;25:1148-1156.
249. Vlaeyen JW, Linton SJ. Fear-avoidance and its consequences in
chronic musculoskeletal pain: A state of the art. Pain 2000;85:317332.
250. Menefee LA, Cohen MJ, Anderson WR, Doghramji K, Frank ED,
Lee H. Sleep disturbance and nonmalignant chronic pain: A
comprehensive review of the literature. Pain Med 2000;1:156172.
251. Bauermeister W. Diagnose und Therapie des Myofaszialen
Triggerpunkt Syndroms durch Lokalisierung und Stimulation
sensibilisierter Nozizeptoren mit fokussierten elektrohydraulische
Stosswellen [German; Diagnosis and therapy of myofascial trigger
point symptoms by localization and stimulation of sensitized
nociceptors with focused ultrasound shockwaves]. MedizinischOrthopädische Technik 2005;5:65-74.
252. Müller-Ehrenberg H, Licht G. Diagnosis and therapy of myofascial
pain syndrome with focused shock waves (ESWT). MedizinischOrthopädische Technik 2005;5:1-6.
Myofascial Trigger Points: An Evidence-Informed Review / 221